首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
During the course of infection, transmitted HIV-1 isolates that initially use CCR5 can acquire the ability to use CXCR4, which is associated with an accelerated progression to AIDS. Although this coreceptor switch is often associated with mutations in the stem of the viral envelope (Env) V3 loop, domains outside V3 can also play a role, and the underlying mechanisms and structural basis for how X4 tropism is acquired remain unknown. In this study we used a V3 truncated R5-tropic Env as a starting point to derive two X4-tropic Envs, termed ΔV3-X4A.c5 and ΔV3-X4B.c7, which took distinct molecular pathways for this change. The ΔV3-X4A.c5 Env clone acquired a 7-amino-acid insertion in V3 that included three positively charged residues, reestablishing an interaction with the CXCR4 extracellular loops (ECLs) and rendering it highly susceptible to the CXCR4 antagonist AMD3100. In contrast, the ΔV3-X4B.c7 Env maintained the V3 truncation but acquired mutations outside V3 that were critical for X4 tropism. In contrast to ΔV3-X4A.c5, ΔV3-X4B.c7 showed increased dependence on the CXCR4 N terminus (NT) and was completely resistant to AMD3100. These results indicate that HIV-1 X4 coreceptor switching can involve (i) V3 loop mutations that establish interactions with the CXCR4 ECLs, and/or (ii) mutations outside V3 that enhance interactions with the CXCR4 NT. The cooperative contributions of CXCR4 NT and ECL interactions with gp120 in acquiring X4 tropism likely impart flexibility on pathways for viral evolution and suggest novel approaches to isolate these interactions for drug discovery.For human immunodeficiency virus type I (HIV-1) to enter a target cell, the gp120 subunit of the viral envelope glycoprotein (Env) must engage CD4 and a coreceptor on the cell surface. Although numerous coreceptors have been identified in vitro, the two most important coreceptors in vivo are the CCR5 (3, 11, 19, 22, 24) and CXCR4 (27) chemokine receptors. HIV-1 variants that can use only CCR5 (R5 viruses) are critical for HIV-1 transmission and predominate during the early stages of infection (86, 90). The importance of CCR5 for HIV-1 transmission is underscored by the fact that individuals bearing a homozygous 32-bp deletion in the CCR5 gene (ccr5-Δ32) are largely resistant to HIV-1 infection (15, 49, 84). Although R5 viruses typically persist into late disease stages, viruses that can use CXCR4, either alone (X4 viruses) or in addition to CCR5 (R5X4 viruses), emerge in approximately 50% of individuals infected with subtype B or D viruses (12, 39, 44). Although not required for disease progression, the appearance of X4 and/or R5X4 viruses is associated with a more rapid depletion of CD4+ cells in peripheral blood and faster progression to AIDS (12, 44, 77, 86). However, it remains unclear whether these viruses are a cause or a consequence of accelerated CD4+ T cell decline (57). The emergence of CXCR4-using viruses has also complicated the use of small-molecule CCR5 antagonists as anti-HIV-therapeutics as these compounds can select for the outgrowth of X4 or R5X4 escape variants (93).Following triggering by CD4, gp120 binds to a coreceptor via two principal interactions: (i) the bridging sheet, a four-stranded antiparallel beta sheet that connects the inner and outer domains of gp120, together with the base of the V3 loop, engages the coreceptor N terminus (NT); and (ii) more distal regions of V3 interact with the coreceptor extracellular loops (ECLs) (13, 14, 36-38, 43, 59, 60, 78, 79, 88). Although both the NT and ECL interactions are important for coreceptor binding and entry, their relative contributions vary among different HIV-1 strains (23). For example, V3 interactions with the ECLs, particularly ECL2, serve a dominant role in CXCR4 utilization (7, 21, 50, 63, 72), while R5 viruses exhibit a more variable use of CCR5 domains, with the NT interaction being particularly important (4, 6, 20, 67, 83). Although V3 is the primary determinant of coreceptor preference (34), it is unclear how specificity for CCR5 and/or CXCR4 is determined, and, in particular, it is unknown how X4 tropism is acquired. Several reports have shown that the emergence of X4 tropism correlates with the acquisition of positively charged residues in the V3 stem (17, 29, 87), particularly at positions 11, 24, and 25 (8, 17, 28, 29, 42, 75), raising the possibility that these mutations directly or indirectly mediate interactions with negatively charged residues in the CXCR4 ECLs. However, Env domains outside V3, including V1/V2 (9, 32, 45, 46, 61, 64, 65, 80, 95) and even gp41 (40), can also contribute to coreceptor switching, and it is unclear mechanistically or structurally how X4 tropism is determined.We previously derived a replication-competent variant of the R5X4 HIV-1 clone R3A that contained a markedly truncated V3 loop (47). This Env was generated by introducing a mutation termed ΔV3(9,9), which deleted the distal 15 amino acids of V3. The ΔV3(9,9) mutation selectively ablated X4 tropism but left R5 tropism intact, consistent with the view that an interaction between the distal half of V3 and the ECLs is critical for CXCR4 usage (7, 21, 43, 50, 59, 60, 63, 72). This V3-truncated virus provided a unique opportunity to address whether CXCR4 utilization could be regained on a background in which this critical V3-ECL interaction had been ablated and, if so, by what mechanism. Here, we characterize two novel X4 variants of R3A ΔV3(9,9) derived by adapting this virus to replicate in CXCR4+ CCR5 SupT1 cells. We show that R3A ΔV3(9,9) could indeed reacquire X4 tropism but through two markedly different mechanisms. One X4 variant, designated ΔV3-X4A, acquired changes in the V3 remnant that reestablished an interaction with the CXCR4 ECLs; the other, ΔV3-X4B, acquired changes outside V3 that engendered interactions with the CXCR4 NT. These divergent evolutionary pathways led to profound differences in sensitivity to the CXCR4 antagonist AMD3100, with ΔV3-X4A showing increased sensitivity relative to R3A and with ΔV3-X4B becoming completely resistant. These findings demonstrate the contributions that interactions with distinct coreceptor regions have in mediating tropism and drug sensitivity and illustrate how HIV''s remarkable evolutionary plasticity in adapting to selection pressures can be exploited to better understand its biological potential.  相似文献   

3.
4.
5.
Cryptococcus neoformans serotype A strains, the major cause of cryptococcosis, are distributed worldwide, while serotype D strains are more concentrated in Central Europe. We have previously shown that deletion of the global regulator TUP1 in serotype D isolates results in a novel peptide-mediated, density-dependent growth phenotype that mimics quorum sensing and is not known to exist in other fungi. Unlike for tup1Δ strains of serotype D, the density-dependent growth phenotype was found to be absent in tup1Δ strains of serotype A which had been derived from several different genetic clusters. The serotype A H99 tup1Δ strain showed less retardation in the growth rate than tup1Δ strains of serotype D, but the mating efficiency was found to be similar in both serotypes. Deletion of TUP1 in the H99 strain resulted in significantly enhanced capsule production and defective melanin formation and also revealed a unique regulatory role of the TUP1 gene in maintaining iron/copper homeostasis. Differential expression of various genes involved in capsule formation and iron/copper homeostasis was observed between the wild-type and tup1Δ H99 strains. Furthermore, the H99 tup1Δ strain displayed pleiotropic effects which included sensitivity to sodium dodecyl sulfate, susceptibility to fluconazole, and attenuated virulence. These results demonstrate that the global regulator TUP1 has pathobiological significance and plays both conserved and distinct roles in serotype A and D strains of C. neoformans.The fungal Tup1 proteins function as global repressors which regulate a large number of genes associated with growth, morphological differentiation, and sexual and asexual reproduction. As a consequence, tup1 mutants are known to display numerous phenotypes (9, 19, 42). The deletion of TUP1 in Candida albicans results in constitutive filamentous growth with no budding yeast cells and is accompanied by loss of virulence (2, 32). In Penicillium marneffei, the only dimorphic species known in the genus Penicillium, deletion of the TUP1 homolog, tupA, confers reduced filamentation and abnormality in yeast morphogenesis (38). In the filamentous fungi Aspergillus nidulans and Neurospora crassa, deletion of the TUP1 homologs, rcoA and rco-1, respectively, severely affects growth and sexual and asexual reproduction (12, 46).Cryptococcus neoformans is a bipolar heterothallic basidiomycetous yeast with two serotypes, A and D, and the function of Tup1 has been studied only for serotype D strains (26, 27). While disruption of TUP1 in strains of serotype D did not affect yeast or hyphal cell morphology, it resulted in mating-type-dependent differences, including temperature-dependent growth, sensitivity to 0.8 M KCl, and expression of genes in several other biological pathways (26). Most importantly, tup1Δ strains displayed a peptide-mediated quorum-sensing-like phenomenon in both mating types of serotype D strains which has not been reported for any other fungal species (27).According to genome sequence data, the serotype A reference strain H99 shares 95% sequence identity with the serotype D reference strain JEC21 (29). However, serotype-specific differences between the two strains have been demonstrated in two major signaling pathways, the pheromone-responsive Cpk1 mitogen-activated protein kinase and cyclic AMP (cAMP) (5, 13, 41, 47). In addition, the high-osmolarity glycerol (HOG) pathway also showed regulatory disparity between the two serotypes (1, 8). Since the regulation of peptide-mediated quorum sensing by TUP1 is reported only for serotype D strains, we sought to determine whether the deletion of TUP1 in serotype A strains would have similar consequences. Surprisingly, we found striking differences in the phenotypes manifested by tup1Δ strains of the two serotypes. We report here the serotype-specific differences in TUP1 regulation between A and D strains and the novel regulatory role of TUP1 in maintaining iron/copper homeostasis in C. neoformans.  相似文献   

6.
Cronobacter spp. are opportunistic food-borne pathogens that are responsible for rare but highly fatal cases of meningitis and necrotizing enterocolitis in neonates. While the operon responsible for yellow pigmentation in Cronobacter sakazakii strain ES5 was described recently, the involvement of additional genes in pigment expression and the influence of pigmentation on the fitness of Cronobacter spp. have not been investigated. Thus, the aim of this study was to identify further genes involved in pigment expression in Cronobacter sakazakii ES5 and to assess the influence of pigmentation on growth and persistence under conditions of environmental stress. A knockout library was created using random transposon mutagenesis. The screening of 9,500 mutants for decreased pigment production identified 30 colorless mutants. The mapping of transposon insertion sites revealed insertions in not only the carotenoid operon but also in various other genes involved in signal transduction, inorganic ions, and energy metabolism. To determine the effect of pigmentation on fitness, colorless mutants (ΔcrtE, ΔcrtX, and ΔcrtY) were compared to the yellow wild type using growth and inactivation experiments, a macrophage assay, and a phenotype array. Among other findings, the colorless mutants grew at significantly increased rates under osmotic stress compared to that of the yellow wild type while showing increased susceptibility to desiccation. Moreover, ΔcrtE and ΔcrtY exhibited increased sensitivity to UVB irradiation.Cronobacter spp. (formerly Enterobacter sakazakii) are opportunistic food-borne pathogens that cause rare but life-threatening cases of meningitis, necrotizing enterocolitis, and septicemia in neonates (7, 30, 39, 40). While the pathogen appears to be ubiquitous, powdered infant formula (PIF) has been implicated as the main source of Cronobacter infection, necessitating effective means of both detecting this organism and preventing contamination in the PIF production environment (14, 26, 40).Although white strains have been observed occasionally, the production of yellow pigment on tryptic soy agar (TSA) is still one of the key discriminative criteria in the identification of presumptive Cronobacter spp. isolates via the ISO/TS 22964 standard protocol (3, 6, 11, 25). Studies of which colorless or cream-white strains of Cronobacter spp. (formerly Enterobacter sakazakii) were identified have reported prevalence rates of 8, 13, and 21.4% (6, 11, 24).The pigment''s carotenogenic nature recently was identified in Cronobacter strain ES5 on a molecular and chemical level (31). Carotenoids are known to stabilize cellular membranes and influence membrane fluidity (13, 22, 48). Functioning as antioxidants, carotenoids scavenge reactive oxygen species (37, 54, 55). Moreover, pigments play a role in the survival of bacteria in harmful environments and have been found to increase the virulence of pathogens such as Staphylococcus aureus and Erwinia chrysanthemi (32, 33, 44, 55). In Cronobacter strain ES5, a gene cluster comprised of seven genes (crtE-idi- crtXYIBZ) was found to be responsible for carotenoid biosynthesis (31). While the study mentioned above identified the operon responsible for carotenoid production, the involvement of other genes in pigment expression has not been investigated.Because no research exists on the influence of pigmentation on the fitness and persistence of Cronobacter spp., the potential implications of failing to detect colorless strains of this organism in the PIF production environment are difficult to assess. Thus, the aim of this study was to further describe the genetic basis of the pigmented phenotype of Cronobacter strain ES5 by isolating and characterizing isogenic white mutants via random transposon mutagenesis and subsequent sequencing, and to identify the impact of pigmentation on persistence and growth under conditions of environmental stress by comparing white mutants to the yellow wild type in a variety of growth and inactivation experiments, a macrophage assay, and a phenotype array.  相似文献   

7.
Herpes simplex virus type 1 (HSV-1)-induced cell fusion is mediated by viral glycoproteins and other membrane proteins expressed on infected cell surfaces. Certain mutations in the carboxyl terminus of HSV-1 glycoprotein B (gB) and in the amino terminus of gK cause extensive virus-induced cell fusion. Although gB is known to be a fusogenic glycoprotein, the mechanism by which gK is involved in virus-induced cell fusion remains elusive. To delineate the amino-terminal domains of gK involved in virus-induced cell fusion, the recombinant viruses gKΔ31-47, gKΔ31-68, and gKΔ31-117, expressing gK carrying in-frame deletions spanning the amino terminus of gK immediately after the gK signal sequence (amino acids [aa] 1 to 30), were constructed. Mutant viruses gKΔ31-47 and gKΔ31-117 exhibited a gK-null (ΔgK) phenotype characterized by the formation of very small viral plaques and up to a 2-log reduction in the production of infectious virus in comparison to that for the parental HSV-1(F) wild-type virus. The gKΔ31-68 mutant virus formed substantially larger plaques and produced 1-log-higher titers than the gKΔ31-47 and gKΔ31-117 mutant virions at low multiplicities of infection. Deletion of 28 aa from the carboxyl terminus of gB (gBΔ28syn) caused extensive virus-induced cell fusion. However, the gBΔ28syn mutation was unable to cause virus-induced cell fusion in the presence of the gKΔ31-68 mutation. Transient expression of a peptide composed of the amino-terminal 82 aa of gK (gKa) produced a glycosylated peptide that was efficiently expressed on cell surfaces only after infection with the HSV-1(F), gKΔ31-68, ΔgK, or UL20-null virus. The gKa peptide complemented the gKΔ31-47 and gKΔ31-68 mutant viruses for infectious-virus production and for gKΔ31-68/gBΔ28syn-mediated cell fusion. These data show that the amino terminus of gK modulates gB-mediated virus-induced cell fusion and virion egress.Herpes simplex virus type 1 (HSV-1) specifies at least 11 virally encoded glycoproteins, as well as several nonglycosylated and lipid-anchored membrane-associated proteins, which serve important functions in virion infectivity and virus spread. Although cell-free enveloped virions can efficiently spread viral infection, virions can also spread by causing cell fusion of adjacent cellular membranes. Virus-induced cell fusion, which is caused by viral glycoproteins expressed on infected cell surfaces, enables transmission of virions from one cell to another, avoiding extracellular spaces and exposure of free virions to neutralizing antibodies (reviewed in reference 56). Most mutations that cause extensive virus-induced cell-to-cell fusion (syncytial or syn mutations) have been mapped to at least four regions of the viral genome: the UL20 gene (5, 42, 44); the UL24 gene (37, 58); the UL27 gene, encoding glycoprotein B (gB) (9, 51); and the UL53 gene, coding for gK (7, 15, 35, 53, 54, 57).Increasing evidence suggests that virus-induced cell fusion is mediated by the concerted action of glycoproteins gD, gB, and gH/gL. Recent studies have shown that gD interacts with both gB and gH/gL (1, 2). Binding of gD to its cognate receptors, including Nectin-1, HVEM, and others (12, 29, 48, 59, 60, 62, 63), is thought to trigger conformation changes in gH/gL and gB that cause fusion of the viral envelope with cellular membranes during virus entry and virus-induced cell fusion (32, 34). Transient coexpression of gB, gD, and gH/gL causes cell-to-cell fusion (49, 68). However, this phenomenon does not accurately model viral fusion, because other viral glycoproteins and membrane proteins known to be important for virus-induced cell fusion are not required (6, 14, 31). Specifically, gK and UL20 were shown to be absolutely required for virus-induced cell fusion (21, 46). Moreover, syncytial mutations within gK (7, 15, 35, 53, 54, 57) or UL20 (5, 42, 44) promote extensive virus-induced cell fusion, and viruses lacking gK enter more slowly than wild-type virus into susceptible cells (25). Furthermore, transient coexpression of gK carrying a syncytial mutation with gB, gD, and gH/gL did not enhance cell fusion, while coexpression of the wild-type gK with gB, gD, and gH/gL inhibited cell fusion (3).Glycoproteins gB and gH are highly conserved across all subfamilies of herpesviruses. gB forms a homotrimeric type I integral membrane protein, which is N glycosylated at multiple sites within the polypeptide. An unusual feature of gB is that syncytial mutations that enhance virus-induced cell fusion are located exclusively in the carboxyl terminus of gB, which is predicted to be located intracellularly (51). Single-amino-acid substitutions within two regions of the intracellular cytoplasmic domain of gB were shown to cause syncytium formation and were designated region I (amino acid [aa] positions 816 and 817) and region II (aa positions 853, 854, and 857) (9, 10, 28, 69). Furthermore, deletion of 28 aa from the carboxyl terminus of gB, disrupting the small predicted alpha-helical domain H17b, causes extensive virus-induced cell fusion as well as extensive glycoprotein-mediated cell fusion in the gB, gD, and gH/gL transient-coexpression system (22, 49, 68). The X-ray structure of the ectodomain of gB has been determined and is predicted to assume at least two major conformations, one of which may be necessary for the fusogenic properties of gB. Therefore, perturbation of the carboxyl terminus of gB may alter the conformation of the amino terminus of gB, thus favoring one of the two predicted conformational structures that causes membrane fusion (34).The UL53 (gK) and UL20 genes encode multipass transmembrane proteins of 338 and 222 aa, respectively, which are conserved in all alphaherpesviruses (15, 42, 55). Both proteins have multiple sites where posttranslational modification can occur; however, only gK is posttranslationally modified by N-linked carbohydrate addition (15, 35, 55). The specific membrane topologies of both gK and UL20 protein (UL20p) have been predicted and experimentally confirmed using epitope tags inserted within predicted intracellular and extracellular domains (18, 21, 44). Syncytial mutations in gK map predominantly within extracellular domains of gK and particularly within the amino-terminal portion of gK (domain I) (18), while syncytial mutations of UL20 are located within the amino terminus of UL20p, shown to be located intracellularly (44). A series of recent studies have shown that HSV-1 gK and UL20 functionally and physically interact and that these interactions are necessary for their coordinate intracellular transport and cell surface expression (16, 18, 21, 26, 45). Specifically, direct protein-protein interactions between the amino terminus of HSV-1 UL20 and gK domain III, both of which are localized intracellularly, were recently demonstrated by two-way coimmunoprecipitation experiments (19).According to the most prevalent model for herpesvirus intracellular morphogenesis, capsids initially assemble within the nuclei and acquire a primary envelope by budding into the perinuclear spaces. Subsequently, these virions lose their envelope through fusion with the outer nuclear lamellae. Within the cytoplasm, tegument proteins associate with the viral nucleocapsid and final envelopment occurs by budding of cytoplasmic capsids into specific trans-Golgi network (TGN)-associated membranes (8, 30, 47, 70). Mature virions traffic to cell surfaces, presumably following the cellular secretory pathway (33, 47, 61). In addition to their significant roles in virus-induced cell fusion, gK and UL20 are required for cytoplasmic virion envelopment. Viruses with deletions in either the gK or the UL20 gene are unable to translocate from the cytoplasm to extracellular spaces and accumulated as unenveloped virions in the cytoplasm (5, 15, 20, 21, 26, 35, 36, 38, 44, 55). Current evidence suggests that the functions of gK and UL20 in cytoplasmic virion envelopment and virus-induced cell fusion are carried out by different, genetically separable domains of UL20p. Specifically, UL20 mutations within the amino and carboxyl termini of UL20p allowed cotransport of gK and UL20p to cell surfaces, virus-induced cell fusion, and TGN localization, while effectively inhibiting cytoplasmic virion envelopment (44, 45).In this paper, we demonstrate that the amino terminus of gK expressed as a free peptide of 82 aa (gKa) is transported to infected cell surfaces by viral proteins other than gK or UL20p and facilitates virus-induced cell fusion caused by syncytial mutations in the carboxyl terminus of gB. Thus, functional domains of gK can be genetically separated, as we have shown previously (44, 45), as well as physically separated into different peptide portions that retain functional activities of gK. These results are consistent with the hypothesis that the amino terminus of gK directly or indirectly interacts with and modulates the fusogenic properties of gB.  相似文献   

8.
9.
10.
11.
During yeast sporulation, a forespore membrane (FSM) initiates at each spindle-pole body and extends to form the spore envelope. We used Schizosaccharomyces pombe to investigate the role of septins during this process. During the prior conjugation of haploid cells, the four vegetatively expressed septins (Spn1, Spn2, Spn3, and Spn4) coassemble at the fusion site and are necessary for its normal morphogenesis. Sporulation involves a different set of four septins (Spn2, Spn5, Spn6, and the atypical Spn7) that does not include the core subunits of the vegetative septin complex. The four sporulation septins form a complex in vitro and colocalize interdependently to a ring-shaped structure along each FSM, and septin mutations result in disoriented FSM extension. The septins and the leading-edge proteins appear to function in parallel to orient FSM extension. Spn2 and Spn7 bind to phosphatidylinositol 4-phosphate [PtdIns(4)P] in vitro, and PtdIns(4)P is enriched in the FSMs, suggesting that septins bind to the FSMs via this lipid. Cells expressing a mutant Spn2 protein unable to bind PtdIns(4)P still form extended septin structures, but these structures fail to associate with the FSMs, which are frequently disoriented. Thus, septins appear to form a scaffold that helps to guide the oriented extension of the FSM.Yeast sporulation is a developmental process that involves multiple, sequential events that need to be tightly coordinated (59, 68). In the fission yeast Schizosaccharomyces pombe, when cells of opposite mating type (h+ and h) are mixed and shifted to conditions of nitrogen starvation, cell fusion and karyogamy occur to form a diploid zygote, which then undergoes premeiotic DNA replication, the two meiotic divisions, formation of the spore envelopes (comprising the plasma membrane and a specialized cell wall), and maturation of the spores (74, 81). At the onset of meiosis II, precursors of the spore envelopes, the forespore membranes (FSMs), are formed by the fusion of vesicles at the cytoplasmic surface of each spindle-pole body (SPB) and then extend to engulf the four nuclear lobes (the nuclear envelope does not break down during meiosis), thus capturing the haploid nuclei, along with associated cytoplasm and organelles, to form the nascent spores (55, 68, 81). How the FSMs recognize and interact with the nuclear envelope, extend in a properly oriented manner, and close to form uniformly sized spherical spores is not understood, and study of this model system should also help to elucidate the more general question of how membranes obtain their shapes in vivo.It has been shown that both the SPB and the vesicle trafficking system play important roles in the formation and development of the FSM and of its counterpart in the budding yeast Saccharomyces cerevisiae, the prospore membrane (PSM). In S. pombe, the SPB changes its shape from a compact dot to a crescent at metaphase of meiosis II (26, 29), and its outer plaque acquires meiosis-specific components such as Spo2, Spo13, and Spo15 (30, 57, 68). This modified outer plaque is required for the initiation of FSM assembly. In S. cerevisiae, it is well established that various secretory (SEC) gene products are required for PSM formation (58, 59). Similarly, proteins presumably involved in the docking and/or fusion of post-Golgi vesicles and organelles in S. pombe, such as the syntaxin-1A Psy1, the SNAP-25 homologue Sec9, and the Rab7 GTPase homologue Ypt7, are also required for proper FSM extension (34, 53, 54). Consistent with this hypothesis, Psy1 disappears from the plasma membrane upon exit from meiosis I and reappears in the nascent FSM.Phosphoinositide-mediated membrane trafficking also contributes to the development of the FSM. Pik3/Vps34 is a phosphatidylinositol 3-kinase whose product is phosphatidylinositol 3-phosphate [PtdIns(3)P] (35, 72). S. pombe cells lacking this protein exhibit defects in various steps of FSM formation, such as aberrant starting positions for extension, disoriented extension and/or failure of closure, and the formation of spore-like bodies near, rather than surrounding, the nuclei, suggesting that Pik3 plays multiple roles during sporulation (61). The targets of PtdIns(3)P during sporulation appear to include two sorting nexins, Vps5 and Vps17, and the FYVE domain-containing protein Sst4/Vps27. vps5Δ and vps17Δ mutant cells share some of the phenotypes of pik3Δ cells (38). sst4Δ cells also share some of the phenotypes of pik3Δ cells but are distinct from vps5Δ and vps17Δ cells, consistent with the hypothesis that Pik3 has multiple roles during sporulation (62).Membrane trafficking processes alone do not seem sufficient to explain how the FSMs and PSMs extend around and engulf the nuclei, suggesting that some other mechanism(s) must regulate and orient FSM/PSM extension. The observation that the FSM is attached to the SPB until formation of the immature spore is complete (68) suggests that the SPB may regulate FSM extension. In addition, the leading edge of the S. cerevisiae PSM is coated with a complex of proteins (the LEPs) that appear to be involved in PSM extension (51, 59). S. pombe Meu14 also localizes to the leading edge of the FSM, and deletion of meu14 causes aberrant FSM formation in addition to a failure in SPB modification (60). However, it has remained unclear whether the SPB- and LEP-based mechanisms are sufficient to account for the formation of closed FSMs and PSMs of proper size and position (relative to the nuclear envelope), and evidence from S. cerevisiae has suggested that the septin proteins may also be involved.The septins are a conserved family of GTP-binding proteins that were first identified in S. cerevisiae by analysis of the cytokinesis-defective cdc3, cdc10, cdc11, and cdc12 mutants (41). Cdc3, Cdc10, Cdc11, and Cdc12 are related to each other in sequence and form an oligomeric complex that localizes to a ring in close apposition to the plasma membrane at the mother-bud neck in vegetative cells (12, 20, 25, 41, 47, 77). The septin ring appears to be filamentous in vivo (12), and indeed, the septins from both yeast (11, 20) and metazoans (31, 36, 69) can form filaments in vitro. The yeast septin ring appears to form a scaffold for the localization and organization of a wide variety of other proteins (8, 22), and it forms a diffusion barrier that constrains movement of membrane proteins through the neck region (7, 8, 73). In metazoan cells, the septins are involved in cytokinesis but are also implicated in a variety of other cellular processes, such as vesicular transport, organization of the actin and microtubule cytoskeletons, and oncogenesis (27, 70).In S. cerevisiae, a fifth septin (Shs1) is also expressed in vegetative cells, but the remaining two septin genes, SPR3 and SPR28, are expressed at detectable levels only during sporulation (15, 17). In addition, at least some of the vegetatively expressed septins are also present in sporulating cells (17, 48), and one of them (Cdc10) is expressed at much higher levels there than in vegetative cells (32). The septins present during sporulation are associated with the PSM (15, 17, 48, 51), and their normal organization there depends on the Gip1-Glc7 protein phosphatase complex (71). However, it has been difficult to gain insight into the precise roles of the septins during sporulation in S. cerevisiae (59), because some septins are essential for viability during vegetative growth, and the viable mutants have only mild phenotypes during sporulation (15, 17), possibly because of functional redundancy among the multiple septins.S. pombe seemed likely to provide a better opportunity for investigating the role of septins during spore formation. There are seven septin genes (spn1+ to spn7+) in this organism (23, 41, 63). Four of these genes (spn1+ to spn4+) are expressed in vegetative cells, and their products form a hetero-oligomeric complex that assembles during cytokinesis into a ring at the division site (2, 3, 10, 76, 79). The septin ring is important for proper targeting of endoglucanases to the division site (44), and septin mutants show a corresponding delay in cell separation (10, 41, 44, 76). However, even the spn1Δ spn2Δ spn3Δ spn4Δ quadruple mutant is viable and grows nearly as rapidly as the wild type (our unpublished results), a circumstance that greatly facilitates studies of the septins'' role during sporulation.spn5+, spn6+, and spn7+ are expressed at detectable levels only during sporulation (1, 45, 78; our unpublished results), and spn2+, like its orthologue CDC10 (see above), is strongly induced (45), but the roles of the S. pombe septins in sporulation have not previously been investigated. In this study, we show that the septins are important for the orientation of FSM extension, suggesting that the septins may have a more general role in dynamic membrane organization and shape determination.  相似文献   

12.
The gammaproteobacterium Xenorhabdus nematophila engages in a mutualistic association with an entomopathogenic nematode and also functions as a pathogen toward different insect hosts. We studied the role of the growth-phase-regulated outer membrane protein OpnS in host interactions. OpnS was shown to be a 16-stranded β-barrel porin. opnS was expressed during growth in insect hemolymph and expression was elevated as the cell density increased. When wild-type and opnS deletion strains were coinjected into insects, the wild-type strain was predominantly recovered from the insect cadaver. Similarly, an opnS-complemented strain outcompeted the ΔopnS strain. Coinjection of the wild-type and ΔopnS strains together with uncolonized nematodes into insects resulted in nematode progeny that were almost exclusively colonized with the wild-type strain. Likewise, nematode progeny recovered after coinjection of a mixture of nematodes carrying either the wild-type or ΔopnS strain were colonized by the wild-type strain. In addition, the ΔopnS strain displayed a competitive growth defect when grown together with the wild-type strain in insect hemolymph but not in defined culture medium. The ΔopnS strain displayed increased sensitivity to antimicrobial compounds, suggesting that deletion of OpnS affected the integrity of the outer membrane. These findings show that the OpnS porin confers a competitive advantage for the growth and/or the survival of X. nematophila in the insect host and provides a new model for studying the biological relevance of differential regulation of porins in a natural host environment.The bacterium Xenorhabdus nematophila forms a mutualistic association with the entomopathogenic nematode Steinernema carpocapsae (2). The nonfeeding infective juvenile form of the nematode (IJ) exists in the soil and carries the bacteria in a specialized receptacle region in the anterior intestine (4, 39). The IJ invades susceptible insect species and enters the hemocoel, where exposure to insect hemolymph stimulates the movement of bacteria down the intestine and out of the anus (36, 39). Together, the nematode and bacteria kill the insect host. X. nematophila not only helps to kill the insect but also promotes bioconversion of host macromolecules and tissues to provide nutrients for nematode reproduction and secretes diverse antimicrobial products to suppress competition for the nutrient resources of the insect cadaver (11, 13, 18, 19, 38). In turn, the nematode vectors X. nematophila to new insect hosts and protects it from the competitive environment of the soil. Colonization of the nematode receptacle is predominantly a monoculture process that is initiated by a single cell followed by bacterial proliferation (24, 39). The level of colonization varies from a few cells to several hundreds per nematode and is higher in nematodes reproducing in insects than on bacterial lawns, suggesting that the insect environment provides additional nutrients for bacterial growth (16, 39).Hydrophilic nutrients and antibiotics passively diffuse across the outer membrane of gram-negative bacteria through general porins and substrate-specific channels (17, 29). The most extensively studied general porins, OmpF and OmpC of Escherichia coli (30), are 16-stranded β-barrel proteins that are reciprocally regulated by changes in external osmolarity (12, 21, 41). Although the flow rate through OmpF is greater than OmpC (28), comparison of the resolved crystal structures does not reveal significant physiochemical differences between the two porins (3). The biological significance of the differential regulation of porins with distinct functional properties remains unclear. The major outer membrane protein of X. nematophila, OpnP, was shown to be produced at high levels in exponentially growing cells and is a homologue of OmpF and OmpC (14). OpnP production was not affected by changes in medium osmolarity, and the flow rate measured for the OpnP porin was more similar to the restrictive porin OmpC than to the more permissive OmpF porin (3). As cells transitioned to stationary phase, de novo synthesis of OpnP decreased, while the synthesis of the outer membrane protein, designated OpnS, increased (15, 22).Porin function and regulation have been studied in both pathogenic and symbiotic bacteria. In Vibrio cholerae two well-studied porins, OmpU and OmpT, that possess distinct functional properties have been shown to be differentially regulated (37). OmpU confers resistance to sodium deoxycholate (DC), a major component of bile, as well as polymixin B, detergents, and antimicrobial peptides, while the expression of OmpT alone sensitizes the cell to DC (26, 33). OmpU was thought to be expressed when V. cholerae colonizes the intestine, suggesting that it was required for host colonization (33); however, subsequent findings indicated that neither OmpU nor OmpT were essential for intestinal colonization (34). Recent findings indicated that OmpU may sense membrane perturbations and activate DegS which in turn modulates σE activity (25, 26). In the symbiotic bacterium Vibrio fischeri the deletion of ompU was shown to reduce the efficiency of colonization of the light organ of the Euprymna scolopes squid and increase sensitivity to bile, antimicrobial peptides, and detergent (1). Interestingly, the ompU strain did not display a competitive defect for colonization in the presence of the wild-type strain.In the present study the growth-phase-regulated outer membrane protein OpnS of X. nematophila was identified as a general porin that conferred a competitive advantage for growth in the insect host. OpnP and OpnS were the only general porins identified in the genome of X. nematophila. The reciprocal expression of OpnP and OpnS suggest that they serve distinct biological roles.  相似文献   

13.
14.
15.
16.
17.
18.
19.
The apicomplexan parasite Toxoplasma gondii expresses type II NADH dehydrogenases (NDH2s) instead of canonical complex I at the inner mitochondrial membrane. These non-proton-pumping enzymes are considered to be promising drug targets due to their absence in mammalian cells. We recently showed by inhibition kinetics that T. gondii NDH2-I is a target of the quinolone-like compound 1-hydroxy-2-dodecyl-4(1H)quinolone (HDQ), which inhibits T. gondii replication in the nanomolar range. In this study, the cationic fluorescent probes Mitotracker and DiOC6(3) (3,3′-dihexyloxacarbocyanine iodine) were used to monitor the influence of HDQ on the mitochondrial inner membrane potential (ΔΨm) in T. gondii. Real-time imaging revealed that nanomolar HDQ concentrations led to a ΔΨm collapse within minutes, which is followed by severe ATP depletions of 30% after 1 h and 70% after 24 h. ΔΨm depolarization was attenuated when substrates for other dehydrogenases that can donate electrons to ubiquinone were added to digitonin-permeabilized cells or when infected cultures were treated with the Fo-ATPase inhibitor oligomycin. A prolonged treatment with sublethal concentrations of HDQ induced differentiation into bradyzoites. This dormant stage is likely to be less dependent on the ΔΨm, since ΔΨm-positive parasites were found at a significantly lower frequency in alkaline-pH-induced bradyzoites than in tachyzoites. Together, our studies reveal that oxidative phosphorylation is essential for maintaining the ATP level in the fast-growing tachyzoite stage and that HDQ interferes with this pathway by inhibiting the electron transport chain at the level of ubiquinone reduction.The apicomplexan parasite Toxoplasma gondii contains a single mitochondrion of an elongated tubular structure (28, 32), which shows several significant metabolic differences from the mammalian counterpart (see references 24 and 33 for review). Although the T. gondii mitochondrion harbors the complete set of enzymes for the tricarboxylic acid cycle (15), it lacks the pyruvate dehydrogenase complex (7, 14, 18), which is typically a central enzyme in carbohydrate metabolism that catalyzes the decarboxylation from pyruvate to acetyl coenzyme A. Other mitochondrial pathways for mitochondrial acetyl coenzyme A generation, such as the 2-methylcitrate cycle, are currently under investigation (33). The T. gondii genome predicts the presence of all components necessary for a respiratory chain. Biochemical evidence for oxidative phosphorylation was provided by extracellular T. gondii tachyzoites that were permeabilized with digitonin (39). However, the overall contribution of oxidative phosphorylation to energy production in relation to other ATP-generating pathways has not been satisfactorily clarified for intracellular T. gondii so far.A fundamental difference of the T. gondii and also the Plasmodium falciparum electron transport chains (ETCs) as opposed to the mammalian ETC is the lack of multisubunit complex I, which couples the transfer of electrons from NADH to ubiquinone with the translocation of protons (6). Instead, P. falciparum expresses one isoform (2) and T. gondii expresses two isoforms (22) of so-called “alternative” or type II NADH dehydrogenases (NDH2s). These single-subunit enzymes do not transport protons across the membrane, and they are, in contrast to the NADH-oxidizing activity of complex I, not rotenone sensitive (21, 27). NDH2s can occur in two topological orientations with respect to the inner mitochondrial membrane. Internal enzymes are facing with their active site toward the mitochondrial matrix and use mitochondrial NAD(P)H as the electron donor, while external enzymes use cytosolic NAD(P)H. Up to now, the orientation of the apicomplexan isoforms is unknown.Due to their absence in the mammalian host, NDH2s were proposed to be promising drug targets against Mycobacterium tuberculosis (40). Their suitability as a drug target in Plasmodium is controversial and has been the subject of discussion (16, 17, 38). Previously, it was demonstrated that low-affinity NDH2 inhibitors in micromolar concentrations were able to inhibit the activity of the P. falciparum NDH2 and led to a collapse of the mitochondrial membrane potential (ΔΨm) (2). The only high-affinity NDH2 inhibitors described so far are 1-hydroxy-2-alkyl-4(1H)quinolones with long alkyl-site chains, for example, 1-hydroxy-2-dodecyl-4(1H)quinolone (HDQ) (C12), which possesses structural similarity to ubiquinone. These compounds were shown to inhibit the activities of Yarrowia lipolytica NDH2 (13) and T. gondii NDH2-I (22) with 50% inhibitory concentrations of between 200 and 300 nM. HDQ was also shown to effectively inhibit T. gondii and P. falciparum replication in nanomolar concentrations in tissue cultures (31).We demonstrate in this study that HDQ treatment in nanomolar concentrations leads to a depolarization of the T. gondii ΔΨm within minutes. The subsequent lack of oxidative phosphorylation leads to a ∼70% reduction of the intracellular ATP level within 24 h. This suggests an indispensable role of NDH2 activity in the maintenance of the ΔΨm and in energy metabolism in the tachyzoite stage.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号